Reservoir Management

You might also like

Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 39

PETROLEUM GEOLOGY

Series 6 Petroleum Technology for the Non-Engineer

C. Reservoir Management

C.1. Rock and Fluid Properties

Porosity

Porosity is defined as the percentage or fraction of void space to bulk volume of rock. If the sedimentary
particles of a rock were of uniform size and packing, as shown in Figure 1 , Figure 2 , and Figure 3 , the
calculation of porosity would be a simple exercise in solid geometry. Of course, actual reservoir rock is a
much more complicated mixture of particles, and its porosity must be measured directly from core samples
or estimated by well log analysis. The proportion and distribution of void space in a reservoir rock can be
modified by the processes of cementation, solution, fracturing, and recrystallization. In reservoir
engineering, primary porosity refers to the void spaces remaining after the sedimentation of particles into
the rock matrix. Secondary porosity is caused by solution channels, fractures, and vugs in the bulk volume
of the matrix, and is developed subsequent to the deposition of the rock. Where both types of porosity exist,
the system is referred to as a dual porosity system. The production sequence of a dual porosity system may
be very different from that of a primary porosity system. Table 1 (below) gives a general range of reservoir
rock matrix porosities. Generally, reservoir rock martix porosities in the lower ranges are of commercial
interest only when a secondary porosity system is present.

Donny’s Library: Reservoir Management 1


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

Rock matrix porosities:

Negligible 0-5
Poor 5-10
Fair  10-15
Good  15-20
Very good 20-25+
Table 1 (above; after Levorsen, 1967)

Usually, the porosity enters into our reservoir calculations in estimating hydrocarbon volume in place in the
reservoir. Its fractional value is multiplied by the bulk volume of the reservoir rock, and by the hydrocarbon
fluid saturation (equal to one minus the water saturation) to determine the hydrocarbon pore volume.

HCPV = V (1-Swc)                     (1)

We are interested not only in porosity, to determine the oil or gas in place at any given time and point in the
reservoir, but also in any porosity (and corresponding permeability) variation across the field, which we
must consider in selecting the best location for development wells. In some cases, the porosity may actually
decrease as the reservoir pressure drops during production, to a degree, that can be significant in its effect
on recovery. This happens when the formation is extremely compressible, especially relative to the fluids
within the pore space. In some loosely consolidated oil sands, such as the Bachaquero field, Venezuela, the
compaction of the reservoir as reservoir fluids are removed accounts for more than 50% of total oil
recovery (Merle et al. 1976). Porosity changes of this type must be taken into account in our reservoir
analysis.

Permeability

Permeability is a measure of the ability of a porous medium to transmit fluids. The oilfield unit of
measurement, the darcy, is named after a nineteenth century French engineer who established that the
velocity of flow through a sand-packed filter was directly proportional to the difference in pressure across
the filter and inversely proportional to the length of the filter (Darcy 1856). His work was ultimately
applied to linear flow through reservoir rock and to radial flow into a wellbore. One darcy is defined as the
permeability that will permit a fluid of one centipoise viscosity to flow at a rate of one cubic centimeter per

Donny’s Library: Reservoir Management 2


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
second through a cross-sectional area of one square centimeter when the pressure gradient is one
atmosphere per centimeter (Muskat 1937). Figure 1 shows a schematic of how this quantity is measured.
Permeability is a property of the porous medium only, and absolute permeability is its value when the
porous medium is saturated with a single phase.

Darcy's definition of permeability


was for a porous medium, which was
100% saturated with the flowing
fluid, water. Hydrocarbon reservoirs
normally have two and perhaps three
phases present: both water and oil; or
water and gas; or water, oil, and gas
sharing the pore space of the rock. It
has been found that having more than
one phase present in the pores
reduces the ability of the rock to
transmit any one of the fluid phases.
For this reason, we define the
effective permeability as the
permeability to one phase when there
is more than one phase present in the
pore space. Its value decreases as the
phases' saturation decreases. There is
an effective permeability value for
each phase present. Usually the
effective permeability is expressed as
a fraction of the absolute
permeability, which is the
permeability at 100% saturation of
the flowing fluid. This ratio of effective to absolute permeability is termed the relative permeability, and is
normally displayed as a set of curves as shown in Figure 2 , for an oil and water system. As would be
expected, the relative permeability to oil decreases as the oil saturation decreases and the water saturation
increases above its irreducible or connate value. Conversely, the relative permeability to water increases,
reaching a maximum when the oil saturation is at its residual saturation. This same general principle applies
to any two- or three-phase system.

We can see how the saturation


changes that accompany production
can affect the ability of the fluids to
flow within the formation. For
example, as the pressure is drawn
down near the wellbore and gas
comes out of solution, the relative
permeability to oil can be
significantly reduced. Gas bubbles
inhibit the flow of oil. Similarly, the
condensation of light oil from a wet
gas, as the pressure drops in the
vicinity of the wellbore, will decrease
the relative permeability to gas and
subsequently the gas flow rate.

Absolute permeability can also vary


throughout the reservoir, depending
on the type of formation and the

Donny’s Library: Reservoir Management 3


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
method of its deposition. The reservoir engineer usually works with the geologist to define such
permeability distributions before beginning major reservoir studies. It is also important to note that a
permeability change can be imposed through drilling operations. The plugging of the pore spaces in the
formation immediately adjacent to the wellbore can reduce permeability and create a "damaged" zone.
Since we know from Darcy's work that permeability relates flow rate and pressure drop, such a damaged
zone will cause a reduction of flow rate for a given pressure drop or require a higher pressure drop (lower
wellbore pressure) to maintain a given flow rate. This adversely affects the efficiency of our entire flowing
system.

Just as the formation near a wellbore may be damaged and the permeability reduced, so too may it be
enhanced. We may acidize and/or fracture the formation, thereby increasing the permeability near the
wellbore and reversing the pressure-flow rate effects of a damaged well.

Table 1 (below) gives some examples of porosity and permeability values for productive fields around the
world.

Reservoir Porosity Permeability

Range % Ave % Range md Ave md

Pennsylvania,

Bradford Sandstone

(Devonian) 2-26 15 0.1-500 50

Southern Arkansas,

Smackover Limestone

(Jurassic) 12.5-21.3 16.9 50-2,000 737

Iran, Massid-i-

Sulaiman field Asmari

Limestone (Jurassic) 2-15 0.0005-0.5

Colorado,

Rangely field Weber

Sandstone (Pennsylvanian) 16 20

Kern Co., California,

Stevens Sand

(Upper Miocene) 15-30 20 10-3,000 140

Oklahoma,

Donny’s Library: Reservoir Management 4


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Wilcox Sand (Ordovician) 8-22 16 79-2497 688

Venezuela Cumarebo

field sandstones (Miocene) 3-39 21.7 1-3397 200-300

Colorado, Niobrara

Chalk (Cretaceous) 33-45 0.3-5.0

Indonesia, Arun

Gas field (Miocene) 7-23 16.2 0-1466

Canada, Norman

Wells field-Kee

Scarp Limestone

(Devonian) 9.8 4

North Sea fields:

Forties 27 400

Brent 23 300

Ninian 20 850

Piper 25 1500

Thistle 25 750

France,

Chateaurenard field,

sandstones (Cretaceous) 30 1000

Rumania,

Suplacu De Barcau field,

Unconsolidated sandstones 32 1700

Nigeria, Meren field,

"G" sands (Miocene) ~30 ~1500

Donny’s Library: Reservoir Management 5


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Table 1 (see heading under "References")

Reservoir Fluid Behavior

Although the composition of reservoir gas is different for each reservoir, certain relationships may be
applied to all gases. The volume of gas in a reservoir will expand as its pressure is reduced and contract as
its temperature is reduced. The net effect is given by the gas formation volume factor, which, by definition,
is the volume in the reservoir (measured in barrels) occupied by 1 standard cu ft (i.e., standard surface
conditions, here 14.7 psia/60 F).

in SI Units:

The Z factor is a dimensionless function of pressure and temperature determined by the gas composition,
which accounts for the compressibility of the gas. This factor is useful because we are interested in the
relationship between pressure in the reservoir and withdrawal of gas in reservoir volumes, and because our
only measurement of that withdrawal is the gas volume measured at the surface. So, it is useful to know
how many reservoir barrels are removed when a certain number of cubic feet at standard conditions are
produced at the surface. For example, 1 barrel of a 0.7 gravity gas in a reservoir at 5000 psia (34,470 kPa)
and 200F (366 K), would occupy 1500 cu ft (42.5 m 3) at standard conditions of temperature and pressure
(60°F and 14.7 psia or 289 K and 101 kPa).

Crude oil behavior under varying conditions of pressure and temperature is (unfortunately) much more
complex than the behavior of a dry gas. Gas consists primarily of methane and perhaps smaller amounts of
the lighter hydrocarbons, while crude oil contains a larger percentage of the heavier hydrocarbons.
Hydrocarbons are grouped into one of several molecular series depending on the relative amounts of carbon
and hydrogen atoms making up their molecules. Paraffins (alkanes), napthenes, and aromatics are the most
common series, with paraffins predominating. Rather than define a reservoir according to its composition,
the reservoir engineer will often give a broader classification range. Table 1 (below) gives one such
classification, with typical ranges for composition, gravity, and gas-oil ratio. These categories are not
rigidly defined and may vary according to local usage. Often the only classification given for a crude oil is
based on its specific gravity (density relative to water), which is relatively easy to measure on a small
sample.

Classification and composition of reservoir hydrocarbons:

GOR Range API Gravity


1. Dry gas (no liquid)
2. Wet gas 1 bbl/100 MCF 50-70
3. Condensate 5 to 100 MCF/bbl  50-70 
4. Volatile oil 3000 CF/bbl 40-50
5. Black oil or dissolved gas systems 100-2500CF/bbl  30-40
6. Heavy oil 0 20-25
7. Tar & bitumen 0

 
Typical Composition

Donny’s Library: Reservoir Management 6


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

C1 C2 C3 C4 C5 C6+
1. Dry gas .90 .05 .03 .01 .01 .01
2. Wet gas 
3. Condensate .75 .08 .04 .03 .02 .08
4. Volatile oil  .6-.65 .08 .05 .04 .03 .2-.15
5. Black oil or dissolved gas systems .44 .04 .04 .03 .02 .43
6. Heavy oil .20 .03 .02 .02 .02 .75
7. Tar & bitumen - - - - - .90 

Table 1 (above)

The API gravity normally used within the oil industry is related to specific gravity by the following:

API = (141.5/S.G. @ 60 F) - 131.5 (1)

Compared to specific gravity, the API scale is expanded and inverted so that a range of specific gravities
from, say, 0.74 to 0.97 becomes a range of API from 60 (very light) to 15 (very heavy). Water has an API
gravity of 10. Table 2 (below) shows the API gravities of several well known crudes. Understanding the
PVT (pressure, volume, temperature) behavior of a crude oil mixture is facilitated by developing a phase
diagram for a particular mixture (Dake 1978). If we take a sample of a lighter paraffinic hydrocarbon
(ethane for example) at a pressure and temperature where it is in the liquid state (point A in Figure 1 ) and
subject it to decreasing pressures at a constant temperature (path A-B), the ethane will vaporize at a certain
point and become a gas (molecules are able to escape from liquid to gas phase). If we repeat this
experiment at many different temperatures, we obtain the vapor pressure line shown in Figure 1 . Above the
vapor pressure line, ethane is in the liquid state and below it is in the gas state.

Various crude gravities:

Donny’s Library: Reservoir Management 7


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

Depth (ft) Depth (m) API


 Field, Country
Safaniyah, Saudi Arabia 5100 1554 27.0 
Prudhoe Bay, U.S.A 8210 2502 26.0
Ninian, U.K. 10,000 3048 35.0
Pembina, Canada 3200-5100 975-1554 32-37.0
Forties, U.K. 8000 2438 37.0
Duri, Indonesia  770 235 22.7
Arun, Indonesia 10,600 3231 54.7
Zakum, Abu Dhabi 9100 2774 39.0
Ghawar, Saudi Arabia 6920 2110 34.0
Lagunillas, Venezuela 3000 914 21.5
Eugene Island 330, U.S.A. 5700 1737 26.0
Jay, U.S.A. 15,470 4715 54.0
Cantarell, Mexico 8528 2599 21.3

Table 2 (from International Petroleum Encyclopedia 1983, PennWell Publishing Co.)

If we duplicate the experiment for another, heavier hydrocarbon, say heptane, the resulting phase diagram
and vapor pressure line will shift and look like that of Figure 2 . As we would expect, the heavier heptane
has a greater tendency to be a liquid at lower temperatures and pressures, than ethane. Both curves end at
the critical point where it is no longer possible to distinguish between liquid or gas.

If we now mix these two


hydrocarbons in a 50-50 mixture, the
phase diagram for the mixture
becomes quite different. It is shown
in Figure 3 . We see that there is now
a two-phase region, where both liquid
and gas coexist. This region exists
because with two components
present, not all of the molecules
change state at a single pressure
along a line of constant temperature.
Naturally occurring hydrocarbon
mixtures are, of course, much more
complex than this two-hydrocarbon
mixture. However, the phase diagram
for each crude oil can be defined in
the laboratory as above.

Donny’s Library: Reservoir Management 8


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Such a phase diagram for a crude oil is shown in Figure 4 . This diagram can be constructed in a manner
similar to what we did above, by decreasing the pressure at a constant temperature, and measuring the
amount of liquid and gas present at
each incremental pressure. If we
begin with a system at point A (
Figure 5 ) and reduce the pressure
while keeping the temperature
constant, we will reach the bubble
point, that point where the first
bubble of gas appears. Below the
bubble point we are in two-phase
region and the relative amount of
liquid present decreases as the
pressure is reduced. Continued
pressure reduction will increase the
amount of gas coming out of solution
until we reach the dew point, that
point where the last drop of liquid
exists.

Beyond this point the hydrocarbon mixture is completely in the gas phase. We can, of course, repeat this
procedure at several different temperatures to obtain the diagram shown in Figure 4 . Similar phase
diagrams are shown for other types of hydrocarbon systems in Figure 6 , Figure 7 , and Figure 8 . Note that
as we learned for heptane, the two-phase envelope moves down and to the right as the percentage of higher
molecular weight components in the hydrocarbon mixture increases. For a dry gas reservoir, the reservoir
temperature is greater than the cricondentherm (marked as CT in Figure 6 , the maximum temperature at
which two phases can coexist for the hydrocarbon mixture, regardless of pressure. If we follow the pressure
depletion line in Figure 7 we can see how production of a dry gas from reservoir conditions (A) to separator
conditions (B), still does not cross the dew point and allow the formation of condensate.

Donny’s Library: Reservoir Management 9


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

Figure 7 shows the phase diagram for a


retrograde condensate reservoir. Here, the
behavior of the mixture is such that a
reduction in reservoir pressure at reservoir
temperature initially causes liquids to
condense in the reservoir. Further pressure
reduction, however, causes some of these
liquids to revaporize. The production of the
lighter gaseous hydrocarbons from the
reservoir mixture will cause the mixture to
become heavier and the entire phase
envelope to shift to the right. This means
that significant amounts of condensate could
be left behind in the reservoir.

Donny’s Library: Reservoir Management 10


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Figure 8 shows a typical phase diagram for
an oil mixture. If the reservoir conditions of
pressure and temperature are at point A,
there is only one phase, liquid oil, in the
reservoir. If we follow the isothermal
pressure reduction line, we can see that after
we cross the bubble-point line, we enter the
region of pressure/temperature conditions
where two phases can exist. Here we have a
volume of liquid oil containing an amount
of dissolved gas commensurate with the
pressure and a volume of liberated, or free,
gas. The volume of liberated gas increases
and that of dissolved gas decreases as we
continue to decrease the pressure. As more
and more light hydrocarbons are liberated
from the oil, the density and viscosity of the
heavier oil phase remaining in the reservoir
increases. In a parallel manner, as more and
more heavier components are vaporized, the
gas phase viscosity and density increases
also. Because the gas is produced more
easily than the oil, the increase in gas-oil ratio as the reservoir pressure is reduced can have severe
consequences on recovery.

There is a formation volume factor for


oil just as there is for gas. Oil will
shrink in volume between reservoir and
surface conditions primarily as a result
of the solution gas evolved. Figure 9
shows schematically how this occurs for
oil and water, and how the gas volume
expands. Figure 10 and Figure 11
shows how the formation volume factor
expressed in bbl/STB (m3/Sm3) for oil
and bbl/SCF (m3/Sm3) change with
decreasing reservoir pressure. For
example, 1.2 to 2.0 reservoir barrels per
stock tank barrel (bbl/STB) is a
common range for oil formation volume
factors. As oil is produced from a
reservoir above the bubble point, the
formation volume factor will increase
slightly as the oil expands. At this point,
no gas is being liberated and the
compressed oil is simply expanding as
volumes are removed. However, when
the bubble point is reached, gas begins
to come out of solution, and a volume of the remaining reservoir oil will have less of a difference between
it and its surface volume. Consequently the formation volume factor will decrease as the pressure is
decreased and more gas is liberated from the oil. In addition, the gas, which is liberated from the oil, will
begin to form a free gas phase in the reservoir. This gas volume will react to the change in pressure with
production. When the reservoir pressure is reduced from 5000 psia to 2000 psia (13,790 kPa), the same
reservoir volume of produced gas that expanded to 1500 cu ft (42.5 m 3) will now only expand to 702 cu ft

Donny’s Library: Reservoir Management 11


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
(19.9 m3) at standard conditions. Consequently, there are less standard cubic feet of gas to each reservoir
barrel of gas produced as the pressure depletes with production.

The phase diagram and PVT properties are different for every mixture of hydrocarbons found in nature. For
a given reservoir fluid, the phase behavior is usually determined by catching a sample at near reservoir
conditions (say, during a drillstem test), subjecting it to pressure and temperature variations in the
laboratory, and measuring the resulting changes in liquid and gas volumes and their individual phase
properties. Where samples are not present, the reservoir engineer often uses published correlations.

Fluid Density, Viscosity & Compressibility

UNDER CONSTRUCTION …!

Wettability & Capillary Pressure

UNDER CONSTRUCTION …!

Exercise 1.

Assume we have an oil reservoir with an area of 100 acres (.4047 km 2) and a thickness of 25 feet (7.62 m).
The porosity is 25% and the water saturation is 20%, as determined by log analysis. We have tested an
exploratory well and determined that the reservoir is at the bubble-point pressure and has an oil formation
volume factor of 1.22 RB/STB. We can be assured of recovering 30% of the oil originally in place, but we
must have about 1.0 x 106 STB (158,987 m3) of reserves to economically develop this field. Should we
develop?

Assume we have an oil reservoir with an area of 100 acres (0.4047 km 2) and a thickness of 25 feet (7.62
m). The porosity is 25% and the water saturation is 20%, as determined by log analysis. We have tested an
exploratory well and determined that the reservoir is at the bubble-point pressure and has an oil formation
volume factor of 1.22 RB/STB. We can be assured of recovering 30% of the oil originally in place, but we

Donny’s Library: Reservoir Management 12


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
must have about 1.0 ´ 106 STB (158,987 m 3) of reserves to economically develop this field. Should we
develop?

SOLUTION: Yes. We know from the equation HCPV = V (1-Swc) that the hydrocarbon pore volume
equals:

HCPV = V (1-Swc)

            = (100) (25) (0.25) (1-0.20)

            = 500 acre-ft

From the conversion table we find that an acre ft equals 7758 barrels (or 1.233 m 3). Thus our original oil in
place is equal to 3,879,000 barrels of reservoir volume (616,500 m3).

Since we expect to recover only 30%, and since our formation volume factor at original conditions is 1.22
RB/STB (1.22 m3 /STD m3), we can calculate reserves of:

Reserves = [3,879,000 RB - 1.22 RB/STB] (0.30)

                 = 954,000 STB (152,000 m3)

Given the degree of accuracy in the reservoir volume, porosity, water saturation, and recovery estimates,
this amount would satisfy our criteria for economic development.

C.2. Natural Drive Mechanisms in Petroleum Reservoirs

Natural Drive Mechanisms

The production of oil and gas is possible only because of potential energy stored in the compressed fluids
and rock of the reservoir or because of energy added to the reservoir. Reservoir energy is released when a
pressure difference is imposed between the wellbore and the reservoir, and a well is produced. While this
pressure differential is maintained, fluids will flow from high- to low-pressure. If the pressure at the
wellbore is sufficient to lift the column of fluid, the well will flow; if not, it must be artificially lifted.
Energy sources in a reservoir vary and this largely determines the efficiency of oil and gas recovery. The
sources of energy for oil production are listed in Table 1 (below) . The dominant type of energy determines
the type of "drive" mechanism attributed to a given reservoir. These mechanisms are listed in Table 2
(below).

Reservoir energy sources:


 
Gas dissolved in oil

Free gas under pressure

· gas reservoir
· oil reservoir w/free gas cap

Donny’s Library: Reservoir Management 13


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Fluid pressure

· hydrostatic-hydrodynamic
· compressed water, gas, oil

Elastically compressed rock

Gravity

Combinations of the above

Table 1 (above)
 
Reservoir drive mechanisms:
Solution-gas drive gas dissolved in oil

Cas-cap drive free gap cap under pressure

Water drive hydrostatic/hydrodynamic pressure and compressed water

Gravity drainage density differences of fluids

Table 2 (above)

A reservoir’s overall means of production is usually a result of some combination of two or more of the
drive mechanisms shown in Table 2. In general, water drive reservoirs have the highest primary oil
recoveries, while solution has drive reservoirs have the lowest. Typical recovery ranges are 12 to 37
percent for solution gas drive (median 20 percent), 15 to 60 percent for has cap drive (median 33 percent),
and 28-84 percent for water drive (median 51 percent).

Solution Gas Drive

Natural gas dissolved in oil will come out of solution and form bubbles, which expand as the fluid pressure
is reduced. This action, somewhat like that occurring in the uncorking of a soda bottle, provides the driving
force in a solution-gas drive reservoir (also called dissolved gas drive, internal gas drive, and depletion
drive). When production is first initiated, compressed oil expands in response to the pressure reduction at
the wellbore. This continues until the bubble-point pressure is reached. At the bubble point, gas bubbles
begin to evolve from solution. With further pressure reduction, the expanding bubbles continue to support
production. This occurs until they reach a critical saturation Ñ the saturation where they join together and
begin to flow as a single gas phase. Above this pressure, the expanding gas bubbles tend to drive the oil out
of the reservoir. Below this pressure, the gas phase, because gas has a much lower viscosity than oil, begins
to flow to the wellbore much more rapidly than the oil. More and more free gas is produced with the crude
oil. This drives the pressure down more rapidly and the finite energy source is rapidly depleted. Ultimately,
the wells cease to flow.

Depending on the geology and rock characteristics, the gas may also migrate to the top of the reservoir and
form a secondary gas cap. This gas cap expands as the pressure is reduced but does not significantly add to
the available energy.

Some energy is also supplied by the expansion of connate water, gas dissolved in connate water, and the
reservoir rock itself. In most cases, this contribution is minor ( Table 1, below) relative to that of the
hydrocarbon expansion.

Donny’s Library: Reservoir Management 14


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Compressibilities of reservoir rock and fluids:

 10-6 volume/volume/psi

Formation rock 3-10

Water 2-4

Oil (above bubble point) 5-100

Gas at 1000 psi (6894.8 kPaManagement of solution-gas drive reservoirs centers on maintaining the
reservoir pressure above that of the critical gas saturation. This is accomplished by reducing oil production
rates to benefit from expansion of compressed oil, water, and rock, by shutting in wells that have begun to
produce excessive gas, and by fluid injection. Secondary recovery in solution-gas drive reservoirs, involves
the injection of gas or water to replace produced volumes and thus maintain reservoir pressure. Figure 1
shows a schematic of a typical solution-gas drive reservoir's characteristic production performance. As we
can see, the gas-oil ratio and oil production rate are fairly stable until the reservoir pressure drops below the
bubble point when the critical gas saturation is reached. Thereafter, the gas-oil ratio begins to climb and the
oil rate drops as increasing amounts of free gas, liberated from the oil, are produced. When the available
solution gas is exhausted, the GOR drops and the oil rate levels off somewhat. The reservoir pressure has
been drastically reduced, however, and most of the reservoir energy has been exhausted.

) 500-1300

Gas at 5000 psi (34,473.8 kPa) 50-200

Table 1 (after Craft and Hamkins, 1959, p. 437)

Management of solution-gas drive reservoirs centers on maintaining the reservoir pressure above that of the
critical gas saturation. This is accomplished by reducing oil production rates to benefit from expansion of
compressed oil, water, and rock, by shutting in wells that have begun to produce excessive gas, and by fluid
injection. Secondary recovery in solution-gas drive reservoirs, involves the injection of gas or water to
replace produced volumes and thus maintain reservoir pressure. Figure 1 shows a schematic of a typical
solution-gas drive reservoir's characteristic production performance. As we can see, the gas-oil ratio and oil
production rate are fairly stable until the reservoir pressure drops below the bubble point when the critical
gas saturation is reached. Thereafter, the gas-oil ratio begins to climb and the oil rate drops as increasing
amounts of free gas, liberated from the oil, are produced. When the available solution gas is exhausted, the
GOR drops and the oil rate levels off somewhat. The reservoir pressure has been drastically reduced,
however, and most of the reservoir energy has been exhausted.

Gas-Cap Drive

A gas cap will be present in a reservoir if there is more gas in a reservoir than can be dissolved in the oil at
original reservoir conditions of temperature and pressure. The gas will migrate to the crest of the structure
and be found as an original free gas cap compressed on top of the oil accumulation. As oil is withdrawn
from the oil zone, the gas cap expands, pushing the underlying oil toward producing wells. The larger the
gas cap volume relative to the oil volume, the lower will be the pressure loss in the gas cap for a given
amount of oil production. As pressure declines, solution-gas drive in the oil zone will also support
production. As the expanding gas cap moves downwards and invades the producing oil zone, wells will
begin to produce increasing amounts of gas and ultimately only gas. Efficient management of a gas-cap
drive reservoir requires that the producing intervals of the oil wells be located where the expanding gas cap

Donny’s Library: Reservoir Management 15


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
will not reach them until the maximum amount of oil has been produced. In addition, a high production rate
may pull or "cone" the less viscous gas downward into the producing interval. For this reason, high GOR
wells are often selectively shut in or recompleted in a lower interval to maintain an even expansion of the
gas cap. Figure 1 shows a schematic of typical gas-cap drive reservoir producing characteristics over time.

As we can see, the reservoir pressure


decline is not as steep as for a solution-
gas drive reservoir. The peaks seen in the
gas-oil ratio curve reflect the sequential
shut-in or recompletion of wells as the
expanding gas cap reaches their
perforations. The oil rate will usually not
decline as steeply as in the case of a
solution-gas drive reservoir. Of course,
the oil rate will be more highly dependent
on the structural placement of wells than
in the solution-gas drive case.

Gas or water injection can be utilized to


maintain pressure in a gas-cap drive
reservoir, just as it is practiced in a
solution-gas drive reservoir. When an
expanding gas cap is coupled with water
influx from an aquifer or from injection,
the combination of displacing fluids
provides an efficient recovery
mechanism.

Water Drive

Often the hydrocarbon accumulation in a reservoir is in contact with a large volume of water-saturated,
permeable reservoir rock. Geologically, the volume of the water reservoir or aquifer may be many times
larger than the volume of the hydrocarbon reservoir, and extend over a large area. In some cases, the rock
may actually outcrop at a point where constant recharging of the aquifer is possible.

As oil and gas are produced from such a system, the water in the aquifer may expand to replace the volume
of oil and gas removed. If oil is produced at a rate that is about equal to the "activity," or response of the
aquifer, the reservoir pressure remains at or near its original value and such a reservoir is said to have an
active water drive. This natural water influx provides a very efficient recovery mechanism, particularly if
the producing wells can be located to drain the maximum area effectively. Water influx can "water out"
wells as the invading water reaches a well and the volume of water produced with the oil increases.
Eventually the fraction of oil produced drops below an economic level and the well must be shut in or
recompleted at a higher interval. Water, like gas, can be "coned" upward into a well, and so production
rates must be monitored to control the amount of coning that takes place. Figure 1 shows a typical
production history of a water drive reservoir.

Donny’s Library: Reservoir Management 16


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
As we can see, the decline in
reservoir pressure with production is
very slight. In fact, for a strong water
drive and low production rates, the
reservoir pressure may remain
essentially unchanged. Because the
pressure does not drop, the producing
GOR remains at or near the initial
solution gas-oil ratio. The oil rate will
remain at a relatively stable level
until the encroaching water begins to
reach the producing wells. As
production continues, the overall
water cut will increase and eventually
all wells may water out.

Gravity Drainage

In addition to solution gas drive, gas


cap drive and water drive, the force of
gravity will cause oil to move
downdip relative to gas, and oil updip
relative to water. Recovery efficiency
is increased if vertical permeability is
great and withdrawal rates are low. Steeply dipping reservoirs are classic examples of where gravity
drainage provides an effective recovery mechanism. The east Texas field is a case in point. Over a 50 year
period, it produced 5 billion barrels (7.95  108 m3) of oil from the steeply dipping Woodbine Formation
by injection of water into downdip wells as they were watered out by the vertical movement of water.

Gas Reservoirs

If the reservoir hydrocarbon composition is such that the accumulation exists as a gas at reservoir
conditions, production of such a gas reservoir is usually a matter of simple gas expansion, sometimes
assisted by water influx. Recovery is generally very good, particularly if the bottomhole flowing pressure
can be reduced to a minimum, by lowering the backpressure at the wellhead with compressors. In such
cases, recovery can be as high as 85% to 90%.

A gas condensate reservoir can be


described as one that produces light
colored, or colorless, liquids with
gravities above 45 API, at GORs in
the range of 5000 to 100,000 SCF/B
(900-18,000 m3/m3) (Craft &
Hawkins 1959). However, these
parameters vary with producing
conditions and measurement
facilities, so the more technical
definition described earlier and
displayed in Figure 1 should be used.

Donny’s Library: Reservoir Management 17


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

Material Balance

If we visualize the reservoir as a "tank" of pressurized fluids, we may observe the changes in reservoir
pressure and producing characteristics as fluids are produced, and thereby determine the type(s) of drive
mechanism(s), in effect, the original volumes available, and the expected recovery. The material balance
equation is the foundation of the reservoir engineer's analysis of a reservoir. Based on the law of
conservation of mass converted to a volume relationship, it is simply expressed as:

volume in-volume out = net change in volume.

This relationship can be restated in terms of reservoir quantities. For a given amount of production and the
associated pressure change, the formula is as follows;

Reservoir Withdrawal = Expansion of oil and originally dissolved gas


 
+ Expansion of gas cap

+ Reduction in hydrocarbon pore volume due to rock and water


expansion

+ Water influx

As a relationship between pressure drop and volume changes, the material balance equation is very
valuable because it allows us to make an estimate of the original volume of hydrocarbons based on the
Pressure-production performance. This estimate is relatively independent of the three-dimensional
geological interpretation. Of course, in order to apply the equation to determine OOIP (original oil in
place), we must have some production and pressure data. We also need some idea of the PVT behavior of
the reservoir fluids, the compressibility of the reservoir rock and the relative size of any gas cap. For a
detailed discussion of material balance relationship, refer to the IPIMS topic "Fundamentals of Petroleum
Reservoir Engineering," which appears in the Reservoir Engineering series in the Petroleum Engineering
Discipline.

C.3. Improve Oil Recovery

Improved Oil Recovery

The use of reservoir energy to produce oil and gas generally results in a recovery of less than 50% of the
original oil in place. The primary recovery mechanisms of solution gas drive, gas cap drive and water drive,
or a combination of one or more of these and gravity drainage, account for most of the world's oil
production. Secondary recovery techniques, in which external energy is added to a reservoir to improve the
efficiency of the primary recovery mechanisms, have been in use for many years. The injection of water to
supplement natural water influx has become an economical and predictable recovery method and is applied
worldwide. Less commonly, gas injection has been used to displace oil downdip in "attic" oil recovery
projects or to maintain gas cap pressure. Still, both primary and secondary recovery techniques have only
been effective in producing roughly one third of the oil discovered. The remaining two thirds, more than
300 billion barrels (4.7696  1010 m3) in the United States alone, is a target for more sophisticated
processes. Such processes, developed to increase recovery from reservoirs considered depleted by primary

Donny’s Library: Reservoir Management 18


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
mechanisms and secondary methods of water or gas injection, were historically termed tertiary recovery
techniques. However, because some of these processes may be applied earlier in the life of a reservoir,
perhaps even in the first day of production, the "tertiary" term is no longer appropriate here, and as a result,
the term enhanced oil recovery methods has been introduced as the term to be used for all processes that
attempt to alter the physical forces that control the movement of oil within the reservoir.

Both conventional water and gas injection, and the more unconventional enhanced oil recovery methods
can collectively be termed improved oil recovery methods.

Waterflooding and Recovery Efficiency

All improved recovery methods involve the injection of fluids into the reservoir via one or more wells, and
the production of oil (and perhaps ultimately the injected fluid) from one or more other wells. The methods
differ in the nature of the fluids used and the physical changes they bring about in the reservoir, but water
usually plays a part in helping to displace the oil. The amount of oil recovered (and obviously the success
of the project) is dependent upon the percentage of oil in place that is contacted and moved by the
displacing fluid. This concept is represented by the equation:

Np = N EV ED                                                               (1)

where : EV = EAS EVS

The oil recovered (Np) is the product of the volume of the oil in place (N), the fraction that is contacted
(EV), and the fraction of the oil contacted that is displaced (E D). The volumetric sweep efficiency (E V),
given as a fraction, is the product of the areal sweep efficiency (E AS), and the vertical sweep efficiency
(EVS). Usually all of these values (except N) increase during the life of an improved recovery project, until
an economic limit is reached. Enhanced recovery methods differ in the manner in which they attempt to
improve either EV or ED. We will discuss each factor separately.

VOLUMETRIC SWEEP EFFICIENCY: The volumetric sweep efficiency (EV), at a given point in time, is
the fraction of the total reservoir volume contacted by the injected fluid during an improved recovery
project. It is the composite of the areal sweep efficiency (EAS), and the vertical sweep efficiency (EVS).

When oil is swept from a reservoir by water, an important factor in determining the areal and vertical
sweep efficiencies is the difference in the mobilities of the two fluids. The mobility of any fluid in a porous
medium such as reservoir rock is directly proportional to its velocity of flow and is equal to the effective
permeability to that fluid divided by the fluid viscosity. For oil this would be equal to k 0/0. Because our
reservoir permeability information is available in terms of relative permeability, the mobility is expressed
as:

                                                  (2)

The mobility ratio is defined as the mobility of the displacing phase in the portion of the reservoir contacted
by the injected fluid, divided by the mobility of the displaced phase in the non-invaded portion of the
reservoir. In the case of water displacing oil (waterflooding):

Donny’s Library: Reservoir Management 19


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
If M is less than or equal to one, it means that the oil is capable of traveling at the same or greater velocity
than the water, under the same conditions. The water, therefore, will not bypass the oil and will instead
push it ahead. If M is greater than one, the water is capable of moving faster than the oil and will bypass
some of the oil, leaving unswept areas behind. We can see that an increase in the viscosity of the oil will
cause the mobility ratio to increase. This is logical, as we can imagine attempting to push a viscous, heavy
oil through a pore system and having the less viscous water "finger" through or around the slow moving oil.
An obvious approach to improving the mobility ratio would be to decrease the difference in oil and water
viscosities, by increasing the water viscosity and/or decreasing the oil viscosity.
We can also imagine that the areal
sweep of water through an oil reservoir
would also depend upon where the
water is injected relative to where the
oil is produced. A wide variety of
flooding patterns have been used in the
oil field and some of these are
reproduced in Figure 1 and Figure 2 .
Laboratory models have enabled
researchers to measure the areal sweep
efficiencies for different mobility
ratio/flood pattern combinations.

For example, if we have wells spaced in a five-spot pattern and are producing from a homogeneous
uniform reservoir, the areal sweep efficiency at the point in time when the displacing phase breaks through
to the producing well has been shown to be about 68% to 72% for a mobility ratio of 1.0. Figure 3 (M=1.0)
shows, in stages, the sweep of a five-spot model as the injected fluid moves to breakthrough. Figure 4
shows how the areal sweep efficiency at breakthrough for this pattern changes with mobility ratio. We are
interested in the sweep efficiency at breakthrough because, generally, little additional oil is recovered by
injecting water after a channel of water flow exists between injector and producer.

Donny’s Library: Reservoir Management 20


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

It should also be pointed out that areal variations in permeability will have a major effect on the ability of
the displacing phase to sweep the reservoir. For this reason, the reservoir engineer now works closely with
the development geologist to define the reservoir environment.

Vertical sweep efficiency (EVS) also depends upon the mobility ratio and, in addition, on the vertical
distribution of permeability within
the reservoir. The laboratory
determined areal sweep efficiencies
mentioned above assume a
homogeneous reservoir. If the
permeability varies vertically, as is
often the case in a real reservoir, an
injected fluid will move through the
reservoir with an irregular vertical
front, moving more rapidly in the
more permeable sections. The sweep
efficiency at breakthrough will
depend on the degree of difference in
permeabilities and on the mobility
ratio. Figure 5 shows how changes in
the mobility ratio can affect vertical
sweep. When the mobility ratio is
greater than 1.0, the displacing phase
has more mobility than the displaced
phase. Thus, as displacing fluid
enters the high permeability zone, the
total resistance to flow decreases in
that zone and the flow in that zone
increases. When breakthrough
eventually occurs, a greater portion of
the lower permeability zone is left unswept. When M is less than 1.0, the channeling effect is less
pronounced.

DISPLACEMENT EFFICIENCY: Unfortunately, simply getting the injected fluid to contact a given
volume of the reservoir does not mean that all the oil in that volume will be displaced. The displacement

Donny’s Library: Reservoir Management 21


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
efficiency refers to the fraction of the oil in place that is swept from a unit volume of the reservoir.
Displacement efficiency is a function of fluid viscosities and the relative permeability characteristics of the
reservoir rock (mobility ratio), of the "wettability" of the rock, and of pore geometry.

Wettability pertains to whether water or oil preferentially "wets" the reservoir rock grains. In an oil-wet
reservoir, oil is preferentially adsorbed on the grain surface. For a water-wet reservoir, the converse is true.
Oil reservoirs range from strongly oil-wet, through intermediate gradations to strongly water-wet. Most are
water-wet. In a water-wet system, the oil exists in the middle of the pore system with water distributed
between it and the rock surfaces. As water displaces oil through the porous medium, some of the oil is left
in globules at the center of the pores. In an oil-wet system, the residual oil clings to the rock grains and is
left as a film on the pore walls. Figure 6 ((a) water-wet rock (b) oil-wet rock) illustrates the difference
between these two extremes.

We can also see the role of wettability in the capillary effects in a reservoir. Figure 7 shows a microscopic
view of a water-wet pore system. As the invading water reaches the intersection of the large diameter and
smaller diameter pore channels, the individual pores "pull" the water phase ahead in a manner similar to
that of water rising in a capillary tube. The advance is greatest in the pore with the smallest radius. Once
the water-oil interface has reached a capillary equilibrium position in a pore, the pressure gradient pushes it
along toward the outlet. The interface reaching the outlet first (probably the smaller capillary) rushes
through the outlet to the next pore and isolates the oil in the other capillary. On a three-dimensional level,
the isolated oil is referred to as ganglia and collectively it makes up the residual oil saturation in the swept
portions of the reservoir. The isolated oil will
not be moved by normal pressure gradients in
the reservoir unless the resistance of the water-
oil interface can be broken down by reducing
the inter-facial tension at the interface.

The relative permeability characteristics of a


reservoir rock and the fluid viscosities are the
properties used to determine the displacement
efficiency. As we mentioned in section 2.1.2,
relative permeability is a measure of the rock's
ability to conduct one fluid when two (or
more) fluids are present. It reflects the

Donny’s Library: Reservoir Management 22


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
composite effect of pore geometry, wettability, fluid distribution, and saturation history on the fluid flow
behavior of the rock-fluid system. Figure 6 shows examples of relative permeability curves for both
strongly water-wet and strongly oil-wet rocks.

The displacement efficiency for a waterflood can be calculated using the fluid viscosities and the water-oil
relative permeability characteristics. The procedure is to construct a fractional flow curve, the fractional
flow of water versus water saturation. If this approach is utilized, it can be shown that the displacement
efficiency at breakthrough is higher the lower the oil viscosity. For a strongly water-wet reservoir with
relative permeability characteristics, as shown in Figure 8 , if the oil viscosity is about 2.0 cp (0.002 Pa s)
or lower, the displacement of oil by the water is "piston-like"; that is, all the recoverable oil in the swept
portion is displaced at breakthrough . Usually the displacement is not so efficient, and the average water
saturation in the invaded portion of the rock increases with continued injection ( Figure 9 ) . Figure 10 and
Figure 11 shows schematically the difference between ideal and non-ideal displacement. The flood front is
shown as a vertical line representing the discontinuity of water saturation behind and ahead of the front.

Donny’s Library: Reservoir Management 23


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

We might find that, based on measurements of relative permeability, fluid viscosities, and permeability
distributions, the expected sweep efficiency factors for a proposed waterflood are:

ED = 0.75; EAS = 0.85; EVS = 0.80

Our overall recovery efficiency is then:

ED EAS EVS = 0.75 x  0.85 x 0.80 = 0.51 or 51 %

and we might expect to recover 51% of the oil in place as a result of water-flooding. This means that only
half the oil can be recovered even when we have a good level of efficiencies. The task of recovering the
balance of the oil is formidable!

Enhanced oil recovery methods attempt to improve these efficiency factors by:

 reducing the mobility ratio by increasing water viscosity;

 reducing the mobility ratio by decreasing oil viscosity;

 altering the interfacial tension of the water-oil interface; and

 improving the relative permeability characteristics.

Table 1 (below) lists the enhanced recovery methods that are of importance today.

Enhanced oil recovery methods:

Chemical Recovery Processes


 

Donny’s Library: Reservoir Management 24


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

 Polymer flooding

 Surfactant-polymer flooding (microemulsion flooding)

 Caustic flooding

Thermal recovery processes

 Steam flooding

 In-situ combustion

Miscible recovery processes

 Miscible hydrocarbon displacement

 Carbon dioxide injection

 Inert gas injection

Table 14 (above)

Enhanced Oil Recovery (EOR)

Enhanced oil recovery processes attempt to alter the physical forces holding oil in a reservoir by improving
the volumetric sweep efficiency through mobility ratio control, and by improving the displacement
efficiency through surface active agents, heat, and miscible displacement. Regardless of laboratory
performance, the value of all enhanced oil recovery techniques ultimately is determined by the degree to
which they are economically applicable in the field. It may be possible to maximize recovery by initiating
an enhanced recovery project early in the life of a field, but the additional early capital cost with delayed
return may make it economically unattractive. Economic applicability is critically dependent on an accurate
geological interpretation of the reservoir. By providing guidance throughout the development of a reservoir,
geologists can help provide for the maximum recovery of the oil they have found.

Chemical Recovery Processes

Several approaches that improve the mobility ratio in a waterflood and reduce the interfacial tension at the
oil-water interface have involved the addition of chemical agents to injected water.

POLYMER FLOODING: An obvious approach to improving the mobility ratio would be to increase the
effective viscosity of the injected water before injection into the reservoir. This can be done by the addition
of long chain molecules, called polymers (the jelly-like component in your tube of shampoo). Although
there are many polymers available for this approach, the most economically attractive examples are
polysaccharides (xanthan gum) and polyacrylamides. The polysaccharides are produced by microbial
action while the polyacrylamides are synthetically produced chemicals with a wide range of molecular
weights and chain lengths. Polyacrylamides are relatively economical and stable, but are sensitive to shear
and salinity. Polysaccharides stand up to salinity and shearing rates but are prone to bacterial and thermal
degradation. A commercial polymer solution is a non-Newtonian fluid. Its behavior is generally
characterized as pseudoplastic; that is, its resistance to flow and its apparent viscosity is lower at low flow
velocities than at high flow velocities. However, at very high velocities, such as those that may exist near
the injection wells, the polymer solution can act as a dilatant fluid and its apparent viscosity will increase.

Donny’s Library: Reservoir Management 25


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
(Where high velocities might cause bypassing, the polymer solution increases viscosity and slows down!)
Figure 1 (comparison of viscosities of two types of polymers at 1000 ppm in 1.0% NaCl at 74  F) shows the
effect of polymer solution concentration on viscosity at a given shear rate. (Normal formation water
viscosity is about .2 to 2 cp.)

In addition to increasing water viscosity and thereby reducing the mobility ratio, polymer flooding also
improves areal and vertical sweep efficiency by reducing the relative permeability to the polymer solution.
Apparently this is accomplished by adsorption of the polymer onto the rock grains, by entrapment of
polymer molecules in pore throats, and by the inability of the polymer-laden solution to enter small pore
channels. Overall, the reduction in permeability allows the preferential filling of the high permeability
streaks or zones in the reservoir with a viscous slug, lowering the velocity of flow and increasing the sweep
of the lower permeability zones.

Polymer flooding does not decrease the residual oil saturation significantly in the swept zone. Its primary
importance is in the improvement of the areal and vertical sweep efficiencies and the acceleration of oil
production before the economic limit of water-oil production ratio is reached. Polymer flooding is most
efficient when begun early in the life of a waterflood, particularly when mobility ratios are poor (2 to 20)
and significant permeability variations exist.

SURFACTANT POLYMER FLOODING: Surface active agents, or surfactants, are compounds that can
act to reduce the interfacial tension at the interface between the oil and water in the reservoir. Detergents
are examples of an everyday use of surfactants to allow water to displace oil (and the accompanying dirt)
from clothing, dishes, etc. Reduction in interfacial tension improves the displacement efficiency of the
flood and reduces the residual oil saturation. Surfactants are usually introduced to a waterflood as
components in a water-oilsurfactant solution. This solution is also called a microemulsion, or a micellar
solution, because of the existance of micelles, aggregates of surfactant molecules surrounding microscopic
oil droplets in water, or microscopic water droplets in oil. A slug of this solution is injected into the
reservoir, usually in a high volume (15% to 60% of reservoir pore volume) with a low concentration, or in a
low volume (3% to 20% of reservoir pore volume) with a high concentration.

In the low concentration case, the reduction in interfacial tension increases oil recovery gradually with the
passage of increasing volumes of surfactant solution. In the high concentration case, the surfactant solution
rapidly displaces water and almost all the oil contacted, but it reverts to a low-concentration flood as the
surfactant slug becomes diluted by formation fluids. Figure 2 shows schematically how a microemulsion
and a polymer solution, followed by water injection, are used in sequence to improve the mobility of the
flood.

Some surfactant solutions have cosurfactants (usually alcohol), electrolytes, or inorganic salts added to
improve performance. Surfactant slug size and composition must be tailored to the specific reservoir rock
and fluid properties. Adsorption of surfactants onto the rock surface can be an important reason for slug
breakdown. The subject of surfactant solutions is complex and their behavior in the reservoirs is not easily
predicted.

CAUSTIC FLOODING: The fact that the addition of sodium hydroxide (caustic) to injection water
improves oil recovery for many reservoirs has been recognized for many years. Although the process
appears simple and relatively inexpensive, the mechanisms involved in displacement of the oil are
complicated. At present, at least five methods are believed to act in the process:

 lowering of interfacial tension by reaction of the caustic with acidic crude oil
components usually found in heavier, viscous crudes, to create surfactants in situ;

 reversal of rock wettability;

Donny’s Library: Reservoir Management 26


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

 emulsification of residual oil and entrapment in small pore throats to reduce water
mobility, and improve areal and vertical sweep efficiency;

 in-situ emulsification and entrainment of residual oil into the flowing caustic water
phase, carrying the oil out of the rock;

 solubilization of rigid films, which form at the oil-water interface, allowing


mobilization of the residual oil.

The most commonly used caustic chemical is sodium hydroxide, although sodium orthosilicate, ammonium
hydroxide, and others have been suggested. The reaction of the reservoir rock with the chemical is an
important factor that must be considered in designing a caustic flood. Crude oil acidity is also important.
Also, the hardness of the injection water must be controlled, or else precipitates of calcium hydroxide may
form, plugging the injection wells. Low concentration caustic solutions (0.5 to 2.0 wt%) appear to offer the
best results for interfacial tension reduction.

Chemically enhanced oil recovery processes act to improve mobility ratio, increase volumetric sweep, and
improve displacement efficiency. Compared with other methods, recovery costs can be somewhat high,
except for caustic flooding. Table 1 (below) shows some generalized reservoir criteria for screening
reservoir candidates.

Screening criteria for chemical processes:

Surfactant/Polymer

Oil properties  Gravity  API >25

Viscosity cp <30

Composition light intermediates desired

Reservoir Oil saturation >30%

characteristics Formation type sandstone preferred

 Net thickness ft (m) >10 (3)

 Average >20

permeability md

 Depth ft (m) <8000 (2400)

 Temperature  F (K) <175 (353)

Polymer Alkaline

Donny’s Library: Reservoir Management 27


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
(caustic)

Oil properties  Gravity  API >25 13-35

 Viscosity cp <150 <200

 Composition not critical some

organic

acids

Reservoir

characteristics  Oil saturation >10% mobile oil above waterflood residual

 Formation type sandstone preferred; sandstone

carbonate preferred

possible

 Net thickness ft (m) not critical not critical

 Average permeability >10 (normally) >20

 Depth ft (m) <9000 (2700) <9000 (2700)

 Temperature  F (k) <200(367) <200 (367)

Table 1 (Taber and Martin, 1983)

Thermal Recovery Processes

Thermal processes attack the problem of an


unfavorable mobility ratio by heating the
reservoir and its fluids, either by adding heat
via steam or hot water, or by generating heat
within the reservoir by burning some of the
oil in the formation. The most important
effect of adding heat is the sharp reduction
in the oil viscosity and the resulting
decrease in the mobility ratio. Figure 1
shows the effect of temperature on oil
viscosity. Note that the viscosity of a heavy,
10API crude oil at 60 F (288.7 K) is about

Donny’s Library: Reservoir Management 28


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
100,000 cp (100 Pa s), but drops to about 10 cp at 360¡F (0.01 Pa s at 455.4 K). Increasing the reservoir
fluid temperature also reduces interfacial tension and increases the relative permeability to oil. It also
vaporizes some of the oil (lighter ends) as it moves forward in the formation where it condenses to form an
"improved" oil.

STEAM FLOODING: Steam flooding consists of the continuous injection of steam into a reservoir with an
injection-production well pattern similar to a waterflood. As the steam moves out into the reservoir away
from the injection well, its temperature drops from heat losses and it begins to condense as hot water (
Figure 2 ). In the steam zone, oil is actually vaporized. In the hot water zone, the oil expands, its viscosity
drops, the residual saturation is lowered, and the relative permeability increased. All of these effects
improve oil recovery.

An alternative to pattern steam flooding is to inject steam into a well, shut it in for a period of time, and
then back-produce the less viscous hot oil and water located near the wellbore.

This cycle is repeated many times


and is termed cyclic steam injection,
"huff and puff," or steam stimulation.
Operators can begin with this
procedure but later must ultimately
convert to a pattern steam flood to
maintain oil production rates.

Steam flooding is currently the


principal enhanced oil recovery
technique. As of 1995, the world’s
largest steamflood project was in the
Duri field in Indonesia, producing
300,000 STB of oil per day. Other
major steamflood areas include the
United States (particularly California)
and Venezuela.

The primary factors limiting the


application of steam injection to oil
reservoirs are depth and thickness;
depth because of the critical pressure
of steam (3202 psia or 22,086 kPa),
and thickness because of excessive
heat loss to underlying and overlying rock formations in the reservoir. Table 1 (below) gives some data
comparing successful steam floods worldwide. One of the California examples, the South Belridge field,
was developed in the 1940s, but by 1952, production of the heavy 13 API, 1600 cp viscosity crude using
normal production methods had peaked. Cyclic steam injection was introduced and increased rates during
the late 1960s, but the effectiveness of successive steam cycles diminished. Continuous steam injection
began in 1969, and within 4 1/2 years, the oil recovery exceeded the total oil recovered during the
preceding 25 years by both primary recovery and cyclic steam stimulation (Van Poollen 1980).

Field- Sand Depth Reservoir h

(ft) Pressure Net Pay

(psig) (ft)

Donny’s Library: Reservoir Management 29


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Kern River, CA 900 35 60

Inglewood, CA 1,000 120 43

Brea B, CA 4,600 100 189

Coalinga, CA 1,500 300 35

Yorba Linda, CA 2,100 200 32

San Ardo Auginac, CA 2,350 250 150

Mount Poso, CA 1,800 100 60

Yorba Linda, CA 650 325

South Beldridge, CA 1,100 180 91

Midway-Sunset, CA 1,600 50 350

Schoonebeck, Holland 2,600 120 83

Slocum, TX 535 110 40

Smackover, AR 2,000 5 20

Tia Juana, Venezuela 1,600 300 125

Winkleman Dome, WY 1,200 210 73

Field- Sand 0 k, kh/0

Oil Vis- Permeability (md-ft/cp)

cosity (cp) (md)

Kern River, CA 4,000 4,000 60

Inglewood, CA 1,200 6,000 220

Brea B, CA 6 70 2,200

Coalinga, CA 100 5,000 1,750

Yorba Linda, CA 600 500 27

San Ardo Auginac, CA 2,000 3,000 225

Mount Poso, CA 280 15,000 3,210

Donny’s Library: Reservoir Management 30


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Yorba Linda, CA 6,400+ 600

South Beldridge, CA 1,600 3,000 170

Midway-Sunset, CA 4,000 4,000 350

Schoonebeck, Holland 180 5,000 2,300

Slocum, TX 1,300 3,500 1,080

Smackover, AR 75 5,000 1,330

Tia Juana, Venezuela 5,000 2,800 70

Winkleman Dome, WY 900 600 50

Field- Sand Oil Content Steam/Oil Oil Steam

(bbl/acre-ft) Ratio Ratio

(bbl/bbl) (bbl/bbl)

Kern River, CA 1,360 4.0 0.25

Inglewood, CA 1,580 2.0 0.50

Brea B, CA 940 4.8 0.21

Coalinga, CA 1,250 2.8 0.36

Yorba Linda, CA 1,070 4.8 0.21

San Ardo Auginac, CA 1,690  

Mount Poso, CA 1,480 4.8 0.21

Yorba Linda, CA   

South Beldridge, CA 1,820 3.6 0.28

Midway-Sunset, CA   

Schoonebeck, Holland 1,980 2.7 0.37

Slocum, TX 1,400 5.6 0.18

Smackover, AR 1.960 3.0 0.33

Donny’s Library: Reservoir Management 31


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Tia Juana, Venezuela 1,660 1.2 0.83

Winkleman Dome, WY 1,450 5.0 0.20

Table 1 (above)

IN-SITU COMBUSTION: Another way of obtaining the beneficial effects of heat in the reservoir is to
generate the heat in situ, or within the reservoir itself. This can be done by injecting oxygen (air) into the
reservoir by using compressors, and then igniting the crude oil-oxygen mixture. Continued injection of air
will cause the burning front or combustion zone to propagate out into the reservoir, heating the oil ahead of
it, and producing steam and hot gases that drive the oil out of the reservoir. There are three basic forms of
in-situ combustion: dry forward combustion, reverse combustion, and wet combustion. Figure 3 shows
schematically how the combustion zone is propagated outward in the direction of the injection in the dry
forward combustion process. Behind the combustion zone the sand is "burned out" and the combustion
zone is the hottest portion of the flood. Immediately ahead of the combustion zone is the coke region,
where the heavier portions of the crude oil are all that remain to be burned, after the heat from the
approaching combustion zone has "cooked" the oil. Ahead of the coke zone are steam and hot water
generated by heat from the combustion zone, light hydrocarbons distilled from the reservoir crude, and a
bank of oil being pushed ahead by the front.

In the reverse combustion process,


the air is injected from the opposite
direction but ignition is at the same
point, now the producing well. This
approach forces the oil to move
through the preheated reservoir,
allowing easier flow of extremely
heavy crudes. However, more of the
lighter portions of the crude are
consumed by the process.

Wet combustion is an attempt to


transfer the heat from the burned out
portion of the reservoir behind the
combustion front in a dry forward
process to the oil ahead of the
burning zone. Water is injected after
the front has been propagated and is
converted into steam in the reservoir
by the hot rock behind the
combustion zone. This partially
quenches the combustion zone and
spreads the generated heat more
evenly through the reservoir.

Important considerations in selection of reservoirs for in-situ combustion and steamflooding processes are
generalized in Table 2 (below) . Vertical sweep in very thick formations is likely to be poor due to
segregation of the steam and combustion gases. In-situ combustion is particularly appropriate when there is
less rock to heat; that is, when the porosity and oil saturations are high. Economic comparisons of in-situ
combustion versus steam injection depend heavily on the cost of fuel to produce steam.

Screening criteria for thermal processes:

Combustion Steamflooding

Donny’s Library: Reservoir Management 32


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Oil properties Gravity API <40 <25
(10-25 normally)

 Viscosity cp <1000 >20

 Composition some asphaltic not critical


components

Reservoir  Oil saturation >40%-50% >40%-50%

characteristics  Formation type sand or sandstone sand or sandstone


with high porosity with high porosity  Net thickness
ft(m) >10(3) >20(6)

 Average
permeability md >100* >200**

 Depth ft (m) >500 (150) >300-5000


(90-1500)

 Temperature >150 (340)


F (K) not critical

*Transmissibility>20 md ft/cp

**Transmissibility>100 md ft/cp

Thermal processes rely on heat primarily to reduce oil viscosity but also to reduce interfacial tension,
improve relative permeability, and vaporize and expand portions of the oil. Both volumetric sweep
efficiency and displacement efficiency are improved.

Miscible Recovery Processes

Washing one's hands with detergent after an oil change will get the worst of it off, but a little gasoline
(while not recommended) will clean up every trace of oil. That's the difference between surfactants and
miscible displacement. Miscibility means that the interface between the displacing and displaced fluids
disappears, so that the oil is dissolved, and the result is 100% displacement efficiency. Of course, a gasoline
flood is absurd from an economic standpoint, but several other methods approach the goal of complete oil
displacement. These methods differ primarily in the type of solvent used to achieve miscibility: refined
hydrocarbons, liquified hydrocarbon gases, carbon dioxide, or inert gases.

MISCIBLE HYDROCARBON DISPLACEMENT: There are three different miscible hydrocarbon


displacement processes: the Miscible Slug Process, the Enriched Gas Process, and the High-Pressure Lean
Gas Process. In the Miscible Slug Process, a propane slug of perhaps 5% of the reservoir pore volume is
injected, followed by natural gas, or gas and water, to drive the solvent through the reservoir. In the
Enriched Gas Process, a slug of hydrocarbon gas with large amounts of C 2-C6 components is injected in
place of the propane slug. The High-Pressure Lean Gas Process substitutes a lean natural gas mixture in
order to cause vaporization of the C 2-C6 components from the oil to the gas, forming a miscible phase at
high-pressure. Figure 1 shows schematically what happens sequentially in a Miscible Slug Process.

Donny’s Library: Reservoir Management 33


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Unfortunately, the solvent becomes con-
cent rated with oil as it moves through the
reservoir and its ability to dissolve the oil
is reduced. Another major problem is the
poor mobility ratio that exists between the
solvent and the gas that follows it.
Alternate injection of gas and water is
used in an attempt to improve this
situation. The lean gas or enriched gas
processes are particularly suited for areas
where there is no ready gas market and
the necessary size of the propane slug is
economically unattractive.

CARBON DIOXIDE FLOODING:


Carbon dioxide is soluable in both oil and
water, and can be miscible with oil,
making it a candidate for use in miscible
floods. Miscibility of carbon dioxide and
oil is pressure dependent, however, and
not all carbon dioxide floods are miscible
processes. In addition to miscibility, other
factors that allow carbon dioxide to improve oil recovery include:

reduction of crude oil viscosity;

swelling of the crude oil;

increased injectivity; and

solution-gas drive effects.

Carbon dioxide flooding can be applied in a variety of ways:


continuous injection of carbon dioxide;

carbon dioxide injection followed by a less expensive gas;

carbon dioxide injection followed by water; and

simultaneous or alternate injection of carbon dioxide and water (called the WAG
process).

Although carbon dioxide is not immediately miscible with crude oil, under the right conditions of
temperature and pressure, the carbon dioxide will extract components from the oil in a manner similar to
that mentioned earlier for lean gas displacement. This mixture forms a miscible front that efficiently
sweeps the oil to the producing wells. The required pressure for miscibility varies with oil composition and
temperature. For example a 30 API oil requires 1200 psi while a 27 API oil requires 4000 psi at 120F.

Pipelines from Colorado to west Texas are now carrying CO 2 to fields that are currently undergoing, or will
soon undergo, CO2 injection.

Donny’s Library: Reservoir Management 34


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
INERT GAS INJECTION: The use of nitrogen, or "flue gas," as a cheaper substitute for carbon dioxide,
or light hydrocarbon mixtures, has been tested in the laboratory and the field. Combustion gases from
boiler flues or gas engine exhausts are primarily nitrogen and carbon dioxide, and have a larger volume
than the gas burned to produce them. Nitrogen does not achieve miscibility as easily as carbon dioxide, but
it can effectively displace the reservoir gas for sale, leaving the inert gas in the reservoir at abandonment.

An inert gas injection project in the Hawkins field, Texas, is designed so that steam boiler exhaust gas is
injected into the gas cap of a water drive reservoir to prevent loss of gas cap pressure. The steam boilers
that produce the flue gas also drive turbines which, in turn, power compressors to inject the inert gas (Van
Poollen 1980).

Table 1 (below) gives some generalized criteria for the selection of candidate reservoirs for miscible
recovery projects. While miscible displacement holds the promise of improving the displacement
efficiency, in most applications the cost of the miscible fluid is high and much of it is not recoverable.
Reservoir heterogeneity is an important factor in all improved recovery projects, but particularly so in
miscible displacement processes where discontinuities can result in the bypassing of costly injectants.

Screening criteria for miscible processes:

Hydrocarbon
Oil properties Gravity API   >35 
Viscosity cp <10 
Composition high % of C2 - C7
Reservoir Oil saturation  >30% PV 
characteristics
Formation type  sandstone or carbonate
Net thickness ft (m) thin unless dipping
Average permeability
not critical
md
>2000 (600) for LPG to >5000 (1500)
Depth ft (m) 
for H.P. Gas 
Temperature F (K) not critical

Nitrogen & Flue Gas Carbon Dioxide


Oil properties >24 
Gravity API  >26 
>35 for N2
Viscosity cp <10 <15
Composition high % of C1 - C7 high % of C5 - C12
Reservoir
Oil saturation >30% PV >30% PV
characteristics
sandstone or
Formation type sandstone or carbonate
carbonate
Net thickness ft (m) thin unless dipping thin unless dipping
Average permeability
not critical not critical
md
Depth ft (m) >4500 (1370) >2000 (6000) 
Temperature F (K) not critical not critical 

Donny’s Library: Reservoir Management 35


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
Table 2 (Taber and Martin, 1983)

Exercise 1.

List eight different EOR methods and the important mechanisms responsible for increasing oil recovery in
each case.

1. Polymer flooding ¾ reduced mobility ratio and reduced relative permeability to the polymer solution,
which results in improved areal and vertical sweep efficiencies.

2. Surfactant-polymer flooding  improves areal and vertical sweep as in polymer flooding, plus, the
surfactant lowers inter-facial tension to improve the displacement efficiency of the flood.

3. Caustic flooding  lowering of interfacial tension, reversal of rock wettability, emulsification and
entrainment of residual oil, and dissolving of oil-water inter facial films all improves displacement
efficiency. Emulsification of oil may also reduce water mobility and thereby improve areal and vertical
sweep efficiency.

4. Steam flooding  improves mobility ratio by reducing oil viscosity. It also reduces interfacial tension,
increases relative permeability to oil and causes the oil to expand or vaporize, improving displacement
efficiency.

5. In-situ combustion  same as steam flooding.

6. Miscible hydrocarbon displacement  dissolves oil to greatly improve displacement efficiency.

7. CO2 injection  reduces crude viscosity to improve volumetric sweep efficiency. It also swells the
crude and may achieve miscibility, improving the displacement efficiency.

8. Inert gas injection  can have some of the beneficial effects of CO2 injection, although pressure
maintenance is usually the most effective application

C.4. Reservoir Simulation

Simulation

The precise behavior of an oil or gas reservoir over a lifetime of producing conditions is never fully
revealed until the reservoir is ultimately depleted. The stream of information that helps to describe the
reservoir begins with the initial geophysical survey and continues through the exploratory drilling,
development drilling, and production phases of the life of the reservoir or field. At some stage in this
process, the petroleum engineer and geologist will be required to estimate the future performance of the
reservoir, based on the available data, in order to determine the optimum development strategy. This point
may come early, say, after the first exploration well is drilled and evaluated, or later, perhaps after primary
recovery is completed and enhanced recovery processes must be considered. Indeed, the estimation
procedure may be repeated and modified many times during the reservoir's life as additional data is
collected. To make these estimates, one must rely on a model, whose behavior closely matches that of the
actual reservoir, and can be used to test and compare development strategies before they are implemented.
This reservoir model is then coupled with an economic model, which takes into account prices,

Donny’s Library: Reservoir Management 36


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
investments, taxes, discount rates, and so on. The two can be used to compare strategies from an economic
perspective and answer questions such as:

 Should the field be developed?

 What is the best location for development wells?

 What are the best completion and production strategies?

 Are additional wells needed?

 Is waterflooding a viable project?

 Should carbon dioxide flooding be implemented?

The reservoir model can be a physical model, as in the case of sand pack studies of waterflood sweep
efficiency, or a mathematical model. Mathematical models can be relatively simple, for example, decline
curves or the material balance equation; or they can be sophisticated, as in numerical simulation.

In numerical simulation, the conventional equations describing conservation of mass, flow behavior, and
pressure-volume-density relationships for the phases present in the reservoir are combined into a system of
partial differential equations. A grid system is designed for the reservoir and each grid point or node is
assigned values of permeability, porosity, fluid content, and pressure. By solving the system of equations at
finite time steps, the changes in pressure and saturation can be determined throughout the reservoir.

The complexity of these equations and the difficulty of their solution as the number of grid lines and time
steps increases, requires the use of numerical approximation techniques. Such a model can be used to
simulate the reservoir's performance for different producing conditions and reservoir descriptions. The
model may be a two-dimensional areal model, a two-dimensional cross-sectional model, or a three-
dimensional model as shown schematically in Figure 1 and Figure 2 . The reservoir engineer selects the
best model, according to the nature of the reservoir and the objectives of the study.

Donny’s Library: Reservoir Management 37


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer
A similar grid pattern ( Figure 3 ) can be developed for models of individual well performance, for
example, to determine the effects of steam stimulation or water and gas coning on an individual well.

In general, the elements of a reservoir simulation study are as follows:

 Setting Up the Study

Defining the study's objectives

Formulating a model

Data Preparation

Collecting rock and fluid data

Reservoir description

Collecting well performance and completion data

Defining producing conditions

History Matching

Adjusting reservoir parameters to match past performance (usually pressure)

Predicting Performance

Running the simulator and analyzing the output

Sensitivity Analysis

Donny’s Library: Reservoir Management 38


PETROLEUM GEOLOGY
Series 6 Petroleum Technology for the Non-Engineer

Identifying critical parameters

Evaluating alternate strategies for development

Presentation of Results

Graphical information

Video graphics

Some important points the reservoir engineer considers when undertaking a reservoir simulation study
include:
· the selection of the least complicated model (largest time steps, minimum number of cells,
fewest runs) that will provide the necessary accuracy;

· use of good judgment to determine which data is most important and will develop as complete a
description as possible of the reservoir;

· selecting the appropriate model based on study objectives and data quality.

References

UNDER CONSTRUCTION … !

Donny’s Library: Reservoir Management 39

You might also like